Localization

Total Page:16

File Type:pdf, Size:1020Kb

Localization Localization Let A be a commutative ring. By a multiplicative set S ⊂ A, we mean a subset such that 1 2 S −1 −1 and if s; t 2 S then st 2 S. We will construct a ring S A and a homomorphism iS : A ! S A that has the following universal property: −1 For all s 2 S, iS(s) is invertible in S A, and for every ring homomorphism φ : A ! B such −1 that φ(s) is invertible in B for all s 2 S, there exists a unique homomorphism φS : S A ! B such −1 that φ = φS ◦ iS : A ! S A ! B. In the definition of multiplicative set, 0 2 S is technically allowed. However, it is very obvious in that case that S−1A = (0), the zero ring, since 0 invertible implies 1 = 0. This is not very interesting. If 1 6= 0 in A, that is, if A 6= (0), and if 0 62 S, then S−1A 6= (0) will follow as well. If A is an integral domain and 0 62 S, then S−1A is a rather obvious subring of the field of fractions of A, namely, the subring of fractions whose denominators are elements of S. But even in this case the main point is not just the construction of S−1A, but rather the connections between ideals in A and S−1A and the many applications in commutative algebra, number theory, and algebraic geometry. The usual argument with universal properties shows that if S−1A exists, it is unique up to isomorphism. The ring S−1A is called a localization of A. This terminology arises from consideration of rings of continuous functions, with values, say, in a field. If A is a ring of functions on a space X and if Y ⊂ X is a subspace, which could be a single point, let S = S(Y ) ⊂ A denote the subset of functions which have no zeros on Y . It is clear that S is a multiplicative set. The quotients a=s, with a 2 A and s 2 S, represent well-defined germs of functions near Y . That is, each such quotient defines a function in some neighborhood of Y , specifically, the neighborhood where s is non-zero. These germs can be added and multiplied and form a ring, whose algebraic properties reflect the properties of X locally, that is, near Y . The ring A might have zero divisors, some elements of S might even be zero divisors, but the ring of germs of functions near Y which can be expressed as quotients a=s, still makes sense. It is perhaps useful to keep this picture in mind as motivation for the general construction below. Before constructing S−1A, we list some examples of multiplicative sets. (i) S is the set of all non-zero divisors in A. If A is an integral domain, this is just the set of all non-zero elements of A, and S−1A will turn out to be the field of fractions of A. (ii) S = A − P , where P ⊂ A is a prime ideal. In this situation, the localization S−1A is often denoted A(P ), and referred to as the localization of A at the prime P . The ring A(P ) is a local ring, that is, a ring with a unique maximal ideal. If A = A(V ) is the affine coordinate ring of a variety over an algebraically closed field K, and if P = I(p) ⊂ A is the maximal ideal corresponding to a point p 2 V , then the localization A(P ) is a ring of germs of K-valued functions defined on open neighborhoods of p 2 V in the Zariski topology. (It is a little harder to interpret the localization A(P ) as a ring of functions if P = I(W ) ⊂ A is a prime ideal corresponding to an irreducible subvariety W ⊂ V .) (iii) It is obvious that arbitrary intersections of multiplicative sets are multiplicative sets. For example, if fPjg is a family of prime ideals in A, then \j(A−Pj) = A−[jPj is a multiplicative set. (iv) If s 2 A is an element which is not nilpotent, then S = fsn j n ≥ 0g is a multiplicative set. Suppose φ : A ! B is a ring homomorphism so that some element s 2 S is a zero divisor in A and φ(s) 2 B is invertible. If as = 0 2 A then obviously φ(a) = 0 2 B, since φ(a)φ(s) = 0. This −1 observation makes it clear that iS : A ! S A will not be injective if any elements of S are zero- divisors in A. A consequence is that the construction of S−1A is somewhat more subtle than, say, 1 the construction of the field of fractions of an integral domain. On the other hand, this observation about consequences of the existence of zero divisors in S motivates the precise definition of S−1A in general, which we now give. S−1A = f[a=s] j a 2 A; s 2 Sg=∼, where [a0=s0] ∼ [a00=s00] if and only if there exists s 2 S with a0s00s = a00s0s 2 A. −1 The homomorphism iS : A ! S A will be defined by iS(a) = [a=1]. The symbol [a=s] will −1 −1 0 0 0 00 0 00 0 00 0 00 name the element iS(a)iS(s) 2 S A. So we certainly want [a =s ] = [a s =s s ] = [a s s=s s s], and [a00=s00] = [a00s0=s00s0] = [a00s0s=s00s0s]. The usual definition of equality of fractions, a0=s0 = a00=s00 if a0s00 = a00s0, would be adequate if no element of S were a zero divisor. The additional factor of s in the definition above allows for the possibility that a0s00 − a00s0 6= 0, but a0s00s − a00s0s = 0 for some s 2 S. Exercise 1 Define addition and multiplication in S−1A, and verify that these operations are well- defined. Acknowledge that you ought to also check all ring axioms, associative laws, distributive laws, identity elements, etc, but don't bother with all that. −1 Exercise 2 Verify the universal property, stated at the outset, for iS : A ! S A. −1 Exercise 3 Characterize the set of all elements t 2 A such that iS(t) is invertible in S A. (This subset of A is also a multiplicative set, called the saturation of S in A. Localizing A with respect to the saturation of S yields the same ring as localizing with respect to S. This can be seen instantly, since the universal property of S−1A implies that this ring also has the required universal property to be the localization with respect to the saturation of S.) Next, we investigate relations between localization and ideals. Suppose I ⊂ A is an ideal with I \ S = ;. Let A∗ = A=I and let S∗ be the image of S in A∗. Then S∗ is a multiplicative set in A∗. The following exercise expresses the commutativity of localization and residue ring formation. One can either divide by an ideal first and then localize, or localize first and then divide by an appropriate ideal. Exercise 4 (S∗)−1A∗ =∼ S−1A=IS−1A. (Although it is easy enough to establish one-to-one ring homomorphisms in both directions, it is more elegant to verify that the ring on the right has the universal property of the ring on the left, hence it is isomorphic to the ring on the left.) If φ : A ! B is any ring homomorphism there are operations known as contraction and extension which relate ideals in B and A. If J ⊂ B is an ideal, the contraction is J c = φ−1(J) ⊂ A, the inverse image of J in A. If I ⊂ A is an ideal, the extension is Ie = φ(I)B, the ideal generated by the image of I in B. Both the contraction and extension operations preserve inclusions, that is, J1 ⊂ J2 c c ce ec implies J1 ⊂ J2, and similarly for extension. There are also obvious inclusions J ⊂ J and I ⊂ I , from which follow trivially the relations J cec = J c and Iece = Ie. The contraction operation has a number of reasonable properties, but very little can be said in general about extension. For example, the following properties of contraction are all easy to check, while the analogues for extension are all false. (i) If J is prime, primary, or radical, then J c is, respectively, prime, primary, or radical. c c c c (ii) rad J = (rad J) and \(Ji ) = (\Ji) . −1 In case the homomorphism is a localization iS : A ! S A, one can say more about the contraction operation, and, in addition, the extension operation has many good properties. −1 ce e Exercise 5 If JS ⊂ S A is an ideal, then JS = JS. If I ⊂ A is an ideal, then I = f[a=s] j a 2 I; s 2 Sg ⊂ S−1A and Iec = fb 2 A j sb 2 I; for some s 2 Sg. 2 Exercise 6 Ie ⊂ S−1A is proper if and only if I \ S = ;. If I ⊂ A is prime or primary and I \ S = ;, then Ie ⊂ S−1A is, respectively, prime or primary. Moreover, in each of these cases, Iec = I ⊂ A. e e −1 Exercise 7 If I ⊂ A is any ideal, then rad I = (rad I) ⊂ S A. If I1 and I2 are ideals in A, e e e −1 then I1 \ I2 = (I1 \ I2) ⊂ S A.
Recommended publications
  • Adjacency Matrices of Zero-Divisor Graphs of Integers Modulo N Matthew Young
    inv lve a journal of mathematics Adjacency matrices of zero-divisor graphs of integers modulo n Matthew Young msp 2015 vol. 8, no. 5 INVOLVE 8:5 (2015) msp dx.doi.org/10.2140/involve.2015.8.753 Adjacency matrices of zero-divisor graphs of integers modulo n Matthew Young (Communicated by Kenneth S. Berenhaut) We study adjacency matrices of zero-divisor graphs of Zn for various n. We find their determinant and rank for all n, develop a method for finding nonzero eigen- values, and use it to find all eigenvalues for the case n p3, where p is a prime D number. We also find upper and lower bounds for the largest eigenvalue for all n. 1. Introduction Let R be a commutative ring with a unity. The notion of a zero-divisor graph of R was pioneered by Beck[1988]. It was later modified by Anderson and Livingston [1999] to be the following. Definition 1.1. The zero-divisor graph .R/ of the ring R is a graph with the set of vertices V .R/ being the set of zero-divisors of R and edges connecting two vertices x; y R if and only if x y 0. 2 D To each (finite) graph , one can associate the adjacency matrix A./ that is a square V ./ V ./ matrix with entries aij 1, if vi is connected with vj , and j j j j D zero otherwise. In this paper we study the adjacency matrices of zero-divisor graphs n .Zn/ of rings Zn of integers modulo n, where n is not prime.
    [Show full text]
  • Group Homomorphisms
    1-17-2018 Group Homomorphisms Here are the operation tables for two groups of order 4: · 1 a a2 + 0 1 2 1 1 a a2 0 0 1 2 a a a2 1 1 1 2 0 a2 a2 1 a 2 2 0 1 There is an obvious sense in which these two groups are “the same”: You can get the second table from the first by replacing 0 with 1, 1 with a, and 2 with a2. When are two groups the same? You might think of saying that two groups are the same if you can get one group’s table from the other by substitution, as above. However, there are problems with this. In the first place, it might be very difficult to check — imagine having to write down a multiplication table for a group of order 256! In the second place, it’s not clear what a “multiplication table” is if a group is infinite. One way to implement a substitution is to use a function. In a sense, a function is a thing which “substitutes” its output for its input. I’ll define what it means for two groups to be “the same” by using certain kinds of functions between groups. These functions are called group homomorphisms; a special kind of homomorphism, called an isomorphism, will be used to define “sameness” for groups. Definition. Let G and H be groups. A homomorphism from G to H is a function f : G → H such that f(x · y)= f(x) · f(y) forall x,y ∈ G.
    [Show full text]
  • Modules and Vector Spaces
    Modules and Vector Spaces R. C. Daileda October 16, 2017 1 Modules Definition 1. A (left) R-module is a triple (R; M; ·) consisting of a ring R, an (additive) abelian group M and a binary operation · : R × M ! M (simply written as r · m = rm) that for all r; s 2 R and m; n 2 M satisfies • r(m + n) = rm + rn ; • (r + s)m = rm + sm ; • r(sm) = (rs)m. If R has unity we also require that 1m = m for all m 2 M. If R = F , a field, we call M a vector space (over F ). N Remark 1. One can show that as a consequence of this definition, the zeros of R and M both \act like zero" relative to the binary operation between R and M, i.e. 0Rm = 0M and r0M = 0M for all r 2 R and m 2 M. H Example 1. Let R be a ring. • R is an R-module using multiplication in R as the binary operation. • Every (additive) abelian group G is a Z-module via n · g = ng for n 2 Z and g 2 G. In fact, this is the only way to make G into a Z-module. Since we must have 1 · g = g for all g 2 G, one can show that n · g = ng for all n 2 Z. Thus there is only one possible Z-module structure on any abelian group. • Rn = R ⊕ R ⊕ · · · ⊕ R is an R-module via | {z } n times r(a1; a2; : : : ; an) = (ra1; ra2; : : : ; ran): • Mn(R) is an R-module via r(aij) = (raij): • R[x] is an R-module via X i X i r aix = raix : i i • Every ideal in R is an R-module.
    [Show full text]
  • Zero Divisor Conjecture for Group Semifields
    Ultra Scientist Vol. 27(3)B, 209-213 (2015). Zero Divisor Conjecture for Group Semifields W.B. VASANTHA KANDASAMY1 1Department of Mathematics, Indian Institute of Technology, Chennai- 600 036 (India) E-mail: [email protected] (Acceptance Date 16th December, 2015) Abstract In this paper the author for the first time has studied the zero divisor conjecture for the group semifields; that is groups over semirings. It is proved that whatever group is taken the group semifield has no zero divisors that is the group semifield is a semifield or a semidivision ring. However the group semiring can have only idempotents and no units. This is in contrast with the group rings for group rings have units and if the group ring has idempotents then it has zero divisors. Key words: Group semirings, group rings semifield, semidivision ring. 1. Introduction study. Here just we recall the zero divisor For notions of group semirings refer8-11. conjecture for group rings and the conditions For more about group rings refer 1-6. under which group rings have zero divisors. Throughout this paper the semirings used are Here rings are assumed to be only fields and semifields. not any ring with zero divisors. It is well known1,2,3 the group ring KG has zero divisors if G is a 2. Zero divisors in group semifields : finite group. Further if G is a torsion free non abelian group and K any field, it is not known In this section zero divisors in group whether KG has zero divisors. In view of the semifields SG of any group G over the 1 problem 28 of .
    [Show full text]
  • The General Linear Group
    18.704 Gabe Cunningham 2/18/05 [email protected] The General Linear Group Definition: Let F be a field. Then the general linear group GLn(F ) is the group of invert- ible n × n matrices with entries in F under matrix multiplication. It is easy to see that GLn(F ) is, in fact, a group: matrix multiplication is associative; the identity element is In, the n × n matrix with 1’s along the main diagonal and 0’s everywhere else; and the matrices are invertible by choice. It’s not immediately clear whether GLn(F ) has infinitely many elements when F does. However, such is the case. Let a ∈ F , a 6= 0. −1 Then a · In is an invertible n × n matrix with inverse a · In. In fact, the set of all such × matrices forms a subgroup of GLn(F ) that is isomorphic to F = F \{0}. It is clear that if F is a finite field, then GLn(F ) has only finitely many elements. An interesting question to ask is how many elements it has. Before addressing that question fully, let’s look at some examples. ∼ × Example 1: Let n = 1. Then GLn(Fq) = Fq , which has q − 1 elements. a b Example 2: Let n = 2; let M = ( c d ). Then for M to be invertible, it is necessary and sufficient that ad 6= bc. If a, b, c, and d are all nonzero, then we can fix a, b, and c arbitrarily, and d can be anything but a−1bc. This gives us (q − 1)3(q − 2) matrices.
    [Show full text]
  • The Structure Theory of Complete Local Rings
    The structure theory of complete local rings Introduction In the study of commutative Noetherian rings, localization at a prime followed by com- pletion at the resulting maximal ideal is a way of life. Many problems, even some that seem \global," can be attacked by first reducing to the local case and then to the complete case. Complete local rings turn out to have extremely good behavior in many respects. A key ingredient in this type of reduction is that when R is local, Rb is local and faithfully flat over R. We shall study the structure of complete local rings. A complete local ring that contains a field always contains a field that maps onto its residue class field: thus, if (R; m; K) contains a field, it contains a field K0 such that the composite map K0 ⊆ R R=m = K is an isomorphism. Then R = K0 ⊕K0 m, and we may identify K with K0. Such a field K0 is called a coefficient field for R. The choice of a coefficient field K0 is not unique in general, although in positive prime characteristic p it is unique if K is perfect, which is a bit surprising. The existence of a coefficient field is a rather hard theorem. Once it is known, one can show that every complete local ring that contains a field is a homomorphic image of a formal power series ring over a field. It is also a module-finite extension of a formal power series ring over a field. This situation is analogous to what is true for finitely generated algebras over a field, where one can make the same statements using polynomial rings instead of formal power series rings.
    [Show full text]
  • Formal Power Series - Wikipedia, the Free Encyclopedia
    Formal power series - Wikipedia, the free encyclopedia http://en.wikipedia.org/wiki/Formal_power_series Formal power series From Wikipedia, the free encyclopedia In mathematics, formal power series are a generalization of polynomials as formal objects, where the number of terms is allowed to be infinite; this implies giving up the possibility to substitute arbitrary values for indeterminates. This perspective contrasts with that of power series, whose variables designate numerical values, and which series therefore only have a definite value if convergence can be established. Formal power series are often used merely to represent the whole collection of their coefficients. In combinatorics, they provide representations of numerical sequences and of multisets, and for instance allow giving concise expressions for recursively defined sequences regardless of whether the recursion can be explicitly solved; this is known as the method of generating functions. Contents 1 Introduction 2 The ring of formal power series 2.1 Definition of the formal power series ring 2.1.1 Ring structure 2.1.2 Topological structure 2.1.3 Alternative topologies 2.2 Universal property 3 Operations on formal power series 3.1 Multiplying series 3.2 Power series raised to powers 3.3 Inverting series 3.4 Dividing series 3.5 Extracting coefficients 3.6 Composition of series 3.6.1 Example 3.7 Composition inverse 3.8 Formal differentiation of series 4 Properties 4.1 Algebraic properties of the formal power series ring 4.2 Topological properties of the formal power series
    [Show full text]
  • The Fundamental Homomorphism Theorem
    Lecture 4.3: The fundamental homomorphism theorem Matthew Macauley Department of Mathematical Sciences Clemson University http://www.math.clemson.edu/~macaule/ Math 4120, Modern Algebra M. Macauley (Clemson) Lecture 4.3: The fundamental homomorphism theorem Math 4120, Modern Algebra 1 / 10 Motivating example (from the previous lecture) Define the homomorphism φ : Q4 ! V4 via φ(i) = v and φ(j) = h. Since Q4 = hi; ji: φ(1) = e ; φ(−1) = φ(i 2) = φ(i)2 = v 2 = e ; φ(k) = φ(ij) = φ(i)φ(j) = vh = r ; φ(−k) = φ(ji) = φ(j)φ(i) = hv = r ; φ(−i) = φ(−1)φ(i) = ev = v ; φ(−j) = φ(−1)φ(j) = eh = h : Let's quotient out by Ker φ = {−1; 1g: 1 i K 1 i iK K iK −1 −i −1 −i Q4 Q4 Q4=K −j −k −j −k jK kK j k jK j k kK Q4 organized by the left cosets of K collapse cosets subgroup K = h−1i are near each other into single nodes Key observation Q4= Ker(φ) =∼ Im(φ). M. Macauley (Clemson) Lecture 4.3: The fundamental homomorphism theorem Math 4120, Modern Algebra 2 / 10 The Fundamental Homomorphism Theorem The following result is one of the central results in group theory. Fundamental homomorphism theorem (FHT) If φ: G ! H is a homomorphism, then Im(φ) =∼ G= Ker(φ). The FHT says that every homomorphism can be decomposed into two steps: (i) quotient out by the kernel, and then (ii) relabel the nodes via φ. G φ Im φ (Ker φ C G) any homomorphism q i quotient remaining isomorphism process (\relabeling") G Ker φ group of cosets M.
    [Show full text]
  • Homomorphisms and Isomorphisms
    Lecture 4.1: Homomorphisms and isomorphisms Matthew Macauley Department of Mathematical Sciences Clemson University http://www.math.clemson.edu/~macaule/ Math 4120, Modern Algebra M. Macauley (Clemson) Lecture 4.1: Homomorphisms and isomorphisms Math 4120, Modern Algebra 1 / 13 Motivation Throughout the course, we've said things like: \This group has the same structure as that group." \This group is isomorphic to that group." However, we've never really spelled out the details about what this means. We will study a special type of function between groups, called a homomorphism. An isomorphism is a special type of homomorphism. The Greek roots \homo" and \morph" together mean \same shape." There are two situations where homomorphisms arise: when one group is a subgroup of another; when one group is a quotient of another. The corresponding homomorphisms are called embeddings and quotient maps. Also in this chapter, we will completely classify all finite abelian groups, and get a taste of a few more advanced topics, such as the the four \isomorphism theorems," commutators subgroups, and automorphisms. M. Macauley (Clemson) Lecture 4.1: Homomorphisms and isomorphisms Math 4120, Modern Algebra 2 / 13 A motivating example Consider the statement: Z3 < D3. Here is a visual: 0 e 0 7! e f 1 7! r 2 2 1 2 7! r r2f rf r2 r The group D3 contains a size-3 cyclic subgroup hri, which is identical to Z3 in structure only. None of the elements of Z3 (namely 0, 1, 2) are actually in D3. When we say Z3 < D3, we really mean is that the structure of Z3 shows up in D3.
    [Show full text]
  • On Finite Semifields with a Weak Nucleus and a Designed
    On finite semifields with a weak nucleus and a designed automorphism group A. Pi~nera-Nicol´as∗ I.F. R´uay Abstract Finite nonassociative division algebras S (i.e., finite semifields) with a weak nucleus N ⊆ S as defined by D.E. Knuth in [10] (i.e., satisfying the conditions (ab)c − a(bc) = (ac)b − a(cb) = c(ab) − (ca)b = 0, for all a; b 2 N; c 2 S) and a prefixed automorphism group are considered. In particular, a construction of semifields of order 64 with weak nucleus F4 and a cyclic automorphism group C5 is introduced. This construction is extended to obtain sporadic finite semifields of orders 256 and 512. Keywords: Finite Semifield; Projective planes; Automorphism Group AMS classification: 12K10, 51E35 1 Introduction A finite semifield S is a finite nonassociative division algebra. These rings play an important role in the context of finite geometries since they coordinatize projective semifield planes [7]. They also have applications on coding theory [4, 8, 6], combinatorics and graph theory [13]. During the last few years, computational efforts in order to clasify some of these objects have been made. For instance, the classification of semifields with 64 elements is completely known, [16], as well as those with 243 elements, [17]. However, many questions are open yet: there is not a classification of semifields with 128 or 256 elements, despite of the powerful computers available nowadays. The knowledge of the structure of these concrete semifields can inspire new general constructions, as suggested in [9]. In this sense, the work of Lavrauw and Sheekey [11] gives examples of constructions of semifields which are neither twisted fields nor two-dimensional over a nucleus starting from the classification of semifields with 64 elements.
    [Show full text]
  • 6. Localization
    52 Andreas Gathmann 6. Localization Localization is a very powerful technique in commutative algebra that often allows to reduce ques- tions on rings and modules to a union of smaller “local” problems. It can easily be motivated both from an algebraic and a geometric point of view, so let us start by explaining the idea behind it in these two settings. Remark 6.1 (Motivation for localization). (a) Algebraic motivation: Let R be a ring which is not a field, i. e. in which not all non-zero elements are units. The algebraic idea of localization is then to make more (or even all) non-zero elements invertible by introducing fractions, in the same way as one passes from the integers Z to the rational numbers Q. Let us have a more precise look at this particular example: in order to construct the rational numbers from the integers we start with R = Z, and let S = Znf0g be the subset of the elements of R that we would like to become invertible. On the set R×S we then consider the equivalence relation (a;s) ∼ (a0;s0) , as0 − a0s = 0 a and denote the equivalence class of a pair (a;s) by s . The set of these “fractions” is then obviously Q, and we can define addition and multiplication on it in the expected way by a a0 as0+a0s a a0 aa0 s + s0 := ss0 and s · s0 := ss0 . (b) Geometric motivation: Now let R = A(X) be the ring of polynomial functions on a variety X. In the same way as in (a) we can ask if it makes sense to consider fractions of such polynomials, i.
    [Show full text]
  • Arxiv:Math/0005288V1 [Math.QA] 31 May 2000 Ilb Setal H Rjciecodnt Igo H Variety
    Mannheimer Manuskripte 254 math/0005288 SINGULAR PROJECTIVE VARIETIES AND QUANTIZATION MARTIN SCHLICHENMAIER Abstract. By the quantization condition compact quantizable K¨ahler mani- folds can be embedded into projective space. In this way they become projec- tive varieties. The quantum Hilbert space of the Berezin-Toeplitz quantization (and of the geometric quantization) is the projective coordinate ring of the embedded manifold. This allows for generalization to the case of singular vari- eties. The set-up is explained in the first part of the contribution. The second part of the contribution is of tutorial nature. Necessary notions, concepts, and results of algebraic geometry appearing in this approach to quantization are explained. In particular, the notions of projective varieties, embeddings, sin- gularities, and quotients appearing in geometric invariant theory are recalled. Contents Introduction 1 1. From quantizable compact K¨ahler manifolds to projective varieties 3 2. Projective varieties 6 2.1. The definition of a projective variety 6 2.2. Embeddings into Projective Space 9 2.3. The projective coordinate ring 12 3. Singularities 14 4. Quotients 17 4.1. Quotients in algebraic geometry 17 4.2. The relation with the symplectic quotient 19 References 20 Introduction arXiv:math/0005288v1 [math.QA] 31 May 2000 Compact K¨ahler manifolds which are quantizable, i.e. which admit a holomor- phic line bundle with curvature form equal to the K¨ahler form (a so called quantum line bundle) are projective algebraic manifolds. This means that with the help of the global holomorphic sections of a suitable tensor power of the quantum line bundle they can be embedded into a projective space of certain dimension.
    [Show full text]